You are on page 1of 8

PHYSICAL/COMPUTATIONAL FRAMEWORK FOR EGS IN SITU FRACTURE

STIMULATION
Justin Pogacnik1, Peter Leary1 and Peter Malin1
1

Institute of Earth Science and Engineering, The University of Auckland, 58 Symonds Street, Auckland, New Zealand
j.pogacnik@auckland.ac.nz

Keywords: Fully coupled finite element model, deformable


porous media, Engineered/Enhanced Geothermal Systems,
heat and mass transfer, permeability enhancement, powerlaw scaling.

(1)

S (k ) ~ const.

Microresistivity and most other geophysical well-log data


have, however, Fourier power-spectra that scale inversely
with wave number by

ABSTRACT
We employ a fully coupled finite element analysis of a
thermal, hydraulic, and mechanical (THM) energy scheme
to simulate stress/strain damage induced in an in situ
poroperm medium stressed by wellbore fluid pressurization
in the medium. Our poroperm medium is characterized by
two empirical constraints: (i) a normally-populated fracturedensity distribution that percolates fluid via long-range
spatially correlated grain-scale fracture connectivity at all
scale; and (ii) a (potentially) long-tailed (lognormal)
permeability distribution associated with percolation
pathways related to normally distributed porosity
distribution F expressed by = 0exp((-0)) as attested
by clastic reservoir well-core poroperm fluctuation
systematics. The degree of fracture connectivity in such a
medium is parameterized by = ratio of standard
deviations of log and distributions. Small values of
describe low degrees of fracture connectivity and hence low
bulk permeability, while large values of describe high
degrees of fracture connectivity giving high bulk
permeability. Wellbore fluid pressurization creates shear
strains in the fracture-heterogeneous poroperm medium,
putatively generating grain-scale fracture damage additional
to the pre-existing grain-scale fracture damage in the
medium. Injecting grain-scale fracture damage can be seen
as creating new fluid flow pathways and increased bulk
permeability via newly created grain-scale fractureconnectivity. Pressure-induced fracture damage injection
thus leads to greater fluid permeability equivalent to
incrementing the value of the fracture-connectivity
parameter . Such wellbore pressurization could be
conducted in interest of flow stimulation of an interwellbore EGS heat exchange volume.
1. INTRODUCTION
Traditionally, the true spatial variations of porosity and
permeability are ignored in favor of their mean values (e.g.,
Sutter et al., 2011). However, previous work indicates that
it is imperative to consider true spatial fluctuations in
poroperm properties. Leary and Walter (2008) show that
observed tight gas well production unpredictability is
traceable to the false assumption that in situ flow in gas
sands is quasi-uniform rather than spatially fluctuating. On
an economic note, Goldstein et al (2011) cite insufficiently
predictable reliability of geothermal reservoir performance
(and in particular, the [un]predictable reliability of EGS
reservoirs) that is traceable to EGS models based on quasiuniform media.
Quasi-uniform media are effectively assumed to have a
white noise Fourier power-spectra in spatial frequency k,

(2)
over about 5 decades of the length scale from 10-2m to
103m. Leary, et al. 2012 offers a detailed description of the
empirical wellbore-based rules given in equations (2)(4b). Equation (2) can be interpreted to mean that in situ
fracture systems are spatially correlated grain-scale fracture
density networks that fluids percolate through on all scale
lengths. The specific well log properties that obey (2)
include: sonic wave speeds, electrical resistivity, soluble
chemical species density, neutron porosity, and mass
density (Leary, 2002).
Clastic rock well-core data show a close spatial fluctuation
relationship between well-core permeability and wellcore porosity , via long-range fracture connectivity:
(3)
where (3) can be interpreted to mean that porosity controls
permeability through interconnectivity of grain-scale
fracture networks that allow fluids to percolate. Leary and
Walter (2008) offer a detailed description of (2) and (3).
Their approach is different because, while the lognormal or
long-tailed trend in permeability and other well data have
been noted for years, i.e., Law (1944) and Bennion &
Griffiths (1966), there have been few reported attempts to
understand the physics underlying long-tailed permeability
distributions and the normal to lognormal transition in
permeability and various ore grade and trace element
distributions.
Leary, et al. 2012 offer a detailed description of the
integration of (3) to yield:
(4a)
where 0 and parameters that can be determined from
specific well-log and well-core data and
(4b)
where sd is the standard deviation of the well properties
over the entire well. Since is a factor in the exponential
term of (4a), it controls the transition of the permeability
distribution from normal to long-tailed. That is, low values
of result in a nearly normal distribution and high values
result in more lognormal-like distributions. Further, has
New Zealand Geothermal Workshop 2012 Proceedings
19 - 21 November 2012
Auckland, New Zealand

been seen to fluctuate from about 2.5 to 30 in clastic rock


systems analyzed by Leary, et al. 2012. The transition to
long-tailed occurs due to an understandable physical reason
in (4a) rather than being a function of arbitrary statistical
parameters that govern lognormal distributions.

Sanford, and Neuzil 2006). Assuming linear isotropic


elastic behavior, the effective stress is calculated as

This work simulates subsurface fluid flow and heat


transport within a 2-D deformable porous medium with
power-law scaling fluctuations in porosity as per (2)
combined with the aforementioned relationship controlling
permeability (3)-(4b). We are most interested in simulating
injection-induced permeability enhancement. Nathenson
(1999) provides an overview of different permeability
enhancement models and compares the results to
Rosemanowes, Cornwall, UK, field data. The models are
formulated in terms of effective stress, however,
mechanical stress is assumed constant and the models
become dependent only upon fluid pressure. In this work,
we examine the inverse power relation from Nathensons
review. The model was originally calibrated by using the
analytical well to well flow solution from De Wiest (1965)
assuming flow through a uniform medium. Injection
induced damage occurs as a result of mechanical
deformation caused by injection pressurization. In this
work, we seek to extend the inverse power relationship to
include mechanical deformation effects.

where

Cladouhos, et al. (2009) propose that sustainable EGS


requires
controlled
shear
induced
permeability
enhancement termed hydroshearing. To that end, we use a
poroperm medium that accounts for natural spatial
fluctuations in permeability and porosity and analyze the
medium for shear stress concentrations. We then assume
that damage occurs in areas of high shear stress
concentrations and effectively increases fracture
connectivity, not fracture density. The methods and results
of these analyses will be presented here.

(8)

is the standard elasticity tensor,

is the

is the thermal strain


mechanical strain vector, and
vector. In the effective stress equation,
(9)
and
(10)

where

is a matrix of spatial derivatives,

is the

displacement vector, m again is related to the identity


matrix, s is the coefficient of thermal expansion of the
is the temperature.
solid matrix, and
2.2 Mass Balance
Consideration of both the fluid and solid mass balance
equations gives:

(11)

2. EQUATIONS AND NUMERICS


where

2.1 Linear Momentum Balance


A suitable first step in acquiring the finite element system
of equations is to consider the linear momentum balance:
(5)
where

refers to the gradient operator,

stress tensor,

is the Cauchy

is the gravity acceleration vector, and

is

the average density of the entire matrix:


,
where

is the density of the solid grains,

the fully saturated case,

is the solid matrix velocity from

the system elastic modulus,


and

is the permeability matrix,

is fluid flow into the system.

2.3 Energy Balance


Lastly, to fully couple thermal effects, we consider the
enthalpy balance equation:

(6)

(12)

is the

density of the fluid (water) surrounding the grains, and is


the porosity defined as the volume of void space per unit
volume. The Cauchy stress can be expressed as a vector and
split into two components, the effective stress component
and a pore pressure component as
,

is the fluid saturation and has been omitted for

(7)

where
is related to the identity tensor, and is the pore
fluid pressure (Lewis and Schrefler 1998; Ingebritsen,

where

is the specific heat of the fluid,

specific heat of the solid grains, and

is the

is related to the

diffusivity of the medium.


The thermal properties of solid are taken to be constant in
space and time (rock thermal properties vary only slightly,
Clauser and Huenges 1995). Fluid density and fluid
viscosity were taken to have the temperature dependence as
prescribed by Likhachev (2003). The temperature
New Zealand Geothermal Workshop 2012 Proceedings
19 - 21 November 2012
Auckland, New Zealand

dependent equation is valid up to about 250C as long as


the fluid does not undergo a phase change. Likhachev also
formulates the pressure dependence of these parameters, but
the effects are small and have been ignored here. The
possibility does exist, however, to incorporate the pressure
dependence.

and Zienkiewicz and Taylor (2000). The final matrix


system of equations takes the form (13):

2.4 The Finite Element Method


Figure 1 depicts an arbitrary 2-D porous domain with a
boundary that is subject to natural and essential boundary
conditions on portions of denoted h and g, respectively.
Natural boundary conditions enforce forces on the system,
such as external tractions, mass fluid flow, or temperature
fluxes. Essential boundary conditions enforce known
degrees of freedom, such as solid displacements, pore fluid
pressures, and temperature. The domain is considered to be
a fully saturated assemblage of rigid solid grains bonded by
weak cements and having a spatially fluctuating density of
fluid-filled void space between grains. Void space fluids
percolate between voids where grain-grain cement bonds
are ruptured by tectonically imposed finite-strain of the
bulk medium. The controlling physical variable is grainscale fracture density. At a critical density of grain-scale
fractures, long-range spatial correlations between grainscale fractures arise, creating the observed power-law
scaling properties of well log spatial fluctuations (Leary
2002; Leary et al. 2012).

where the unknowns are the nodal solid displacements (u),


pore fluid pressures (p), and temperatures (T). This system
can be re-written as

(14)
and is solved using a single-step finite difference operator
in the standard way.
In this paper, we consider 2D sections of earth in two
perpendicular directions to approximate well to well flow
scenarios. In general, the boundary pressure is equivalent to
the pressure and depth. Roller boundary conditions are
places on the outside walls of the domain that constrain
deformation in the normal direction. An injection well
incurs over pressurization and receiver wells are held
constant at the depth pressure. In the present case, the fluid
temperatures are everywhere at ambient rock temperature;
omitting thermal effects simplifies the interpretation of
computed strain. The material parameters specified in the
model that were used in this paper are given in Table 1.
Table 1. Material Input Parameters

Figure 1. An arbitrary porous domain in a poroperm


medium comprising void fluids (white) and solid grains
(dark) that may or may not be bonded by intact
cements. Void-to-void percolation of fluid is assumed to
occur where grain cements are naturally disrupted by
finite-strain dislocations due to tectonic processes, or in
the present example by high pressure fluids introduced
via a wellbore in the medium. The presence/absence of
cement bonds/disruption is assumed to be highly
spatially erratic and heterogeneous at all scale lengths
throughout the medium; there is no assumed effective
medium or continuum approximation.
The solid-fluid matrix is non-isothermal and grain
assemblages are subject to small strains on all scale lengths.
Therefore, the problem is evaluated as a coupled problem
involving temperature and fluid flow through a deformable
porous medium.
The details of the finite element discretization used here can
be found in Pogacnik (2012) and are adapted from the wellestablished procedures found in Lewis and Schrefler (1998)

3. PERMEABILITY ENHANCEMENT
Nathenson (1999) provides an overview of a few different
permeability enhancement relationships (inverse power,
cubic power, cubic-log, and exponential) based on effective
stress ( - p). He uses a simple 1-D well to well flow
analytical solution with optimized material parameters to
match field data taken from the Rosemanowes, UK,
geothermal field. The enhancement relation favored in
Nathensons analysis was the inverse power relationship
given by:

(15)
New Zealand Geothermal Workshop 2012 Proceedings
19 - 21 November 2012
Auckland, New Zealand

where 0 is baseline permeability, c is the confining stress


(taken to be the horizontal confining stress), is the
Cauchy stress, and p is the pressure. Nathenson states that it
is reasonable to assume = c at depth. Therefore, the
expression takes the form:

concentration. The shear strain concentrations are caused by


injection well overpressurization resulting in fluid
percolating through natural pathways in the medium. We
propose that shear strain concentrations result in an increase
in existing fracture connectivity, , from equation (4). For a
first approximation, we take the isotropic permeability to
increase with increased fracture connectivity following (4)
as:

(16)
This form of the inverse power relation has been
implemented in our finite element code in 2-D where c is
taken to be a constant that is equal in either coordinate
direction x or y.
In order to extend Nathensons work to a more general case
in higher dimsions, we formulate equation (15) in terms of
the full stress tensor with potentially variable initial
principle stress components in the in situ stress state, which
would now be expressed as a second order tensor. A more
general form of equation (15) could be written as

(17)

(22)

where * is the fracture connectivity stimulation factor, xy


is the shear strain at a point,

is the maximum shear

strain in the entire domain, and 0 is the minimum pososity


(taken to be fracture porosity) in the domain. Therefore the
shear strain at a point is normalized by the maximum shear
strain in the domain. This is a first order approximation to a
method for enhanced permeability through shear strain
induced damage. Normalizing the shear strains helps to
visualize areas of shear strain concentration in the domain.
Validating and refining the model is the subject of future
work.

where, in 2-D,
4. NUMERICAL RESULTS AND DISCUSSION
4.1 Uniform Media
(18)

(19)

For the inverse power relation, the same material


parameters presented by Nathenson were used here and are
presented in Table 2.
Table 2. Inverse power material parameters from
Nathenson (1999)

(20)
where h and v are the horizontal and vertical components
of the in situ stress state. h is taken to be the same thing as
c in equation (15). Nathensons inverse power relation
obviously breaks down when the confining stress and
pressure are equal since the permeability approaches
infinity. However, this case could correspond to a
hydrofracking scenario and represents a change in failure
mode from shear failure to normal crack opening. This is
exactly the case that should be avoided in EGS since
hydrofractured super-hiways can allow cold fluid to pass
from well to well.
In order to further ellucidate the effects that mechanical
deformation may have on the permeability in Nathensons
formulation, a third case was set up in which only
mechanical stress effects were considered:

Nathenson provides data from Rosemanowes wells RH12


(injection) and RH15 (receiver) with which to compare
permeability enhancement relations. The general set up is
given by two parallel wells that are 150 m apart in
horizontal distance with an average of 275 m of open-hole
well each. Nathenson states that the open-hole portions of
the wells are offset by 400 m of vertical distance and scales
permeability by an unspecified adjustment. He explains
that the adjustment affects the baseline permeability (0)
but not the function dependence of permeability on
pressure. Figure 2 (left) displays the finite element mesh
used to represent the Rosemanowes scenario. Figure 2
(right) displays the enhanced permeability repsonse using
equation (16) and Figure 3 (left) displays the stress state
between the wells and (right) displays the enhanced
permeability as calculated by equation (21).

(21)
In our analyses, it has become apparent that geometry and
boundary conditions play a significant role in the stress
state reponse of the medium. In addition to the above
formulations given in equations (16), (17), and (21), a
fourth permeability enhancement option was considered
where permeability is increased in areas of shear strain
New Zealand Geothermal Workshop 2012 Proceedings
19 - 21 November 2012
Auckland, New Zealand

Figure 2. Finite element mesh with the pressure result


(Left) and the enhanced permeability result (Right) to
represent the Rosemanowes site data constructed from
specifications found in Nathenson (1999). The left side of
the mesh represents the injection well and the right side
the receiver well. The pressure differential between the
wells is 40 bars. The horizontal distance between the
wells is 150 m and the open-hole height is 275 m. The
initial permeability was uniformly set to 1.31E-15 m2.
Permeability was enhanced by the pressure-dependent
equation (16).

Figure 4. Rosemanowes average flow data compared to


simulation results for no permeabilty enhancement and
the different forms of permeability enhancement
derived from the inverse power law presented in
Nathenson (1999). Permeability enhancement results in
higher fluid velocities and higher flowrates at equivalent
pressure drops between wells.

Figure 3. Response for horizontal component of stress


tensor, xx (left), and enhanced permeability, xx (right),
for permeability enhancement controlled by equation
(21). The areas of compression result in an effective
reduction in permeability, while tensile stress results in
permeability enhancement.
Permeability enhancement ultimately affects the percolation
flow through a porous medium. Figure 4 shows the average
flow results from the Rosemanowes wells and the results
for each different use of the inverse power permeability
enhancement equation.
Figure 5 displays the flow results for a vertical slice through
three-spot horizontal well bore set up. The center well is
overpressurized to 8 MPa over the initial hydrostatic
pressure. The two receiver wells are situated on either side
of the injector well and are fixed at hydrostatic pressure.
The figure also displays the horizontal component of the
mechanical stress tensor (middle) and the final enhanced
permeability result (bottom).

Figure 5. Three-spot horizontal well bore simulation


result for pressure in Pa (top), horizontal component of
mechanical stress in Pa (middle), and horizontal
component of enhanced permeability in m2 (bottom).
Enhanced permeability was computed using equation
(17) and is a function of effective stress. The initial
permeability was uniformly set to 1.31E-15 m2.
4.2 Poroperm Medium
The previous section relies on the assumption that initial
permeability is uniform in the medium. A poroperm
medium was constructed that adheres to the empirical
constraints of equations (2), (3), and (4). Figure (6) displays
the poroperm medium with spatially fluctuating porosity
and permeability values. Permeability enhancement occurs
as a result of injection pressure, stress state, and/or shear
strain induced damage. So far, only injection over pressure

New Zealand Geothermal Workshop 2012 Proceedings


19 - 21 November 2012
Auckland, New Zealand

values of 2 MPa were used in the investigation of


permeability enhancement in the poroperm medium.

Figure 6. Poroperm medium used with realistic spatial


fluctuations in permeabilty in m2 (left) and porosity
(right). Small changes in porosity result in large changes
in permeability. The total flucutation in porosity is 0.13
but permeability fluctuates nearly 2 orders of
magnitude.
Figure 7 displays the well to well flow geometry where
injection stimulation was induced by shear strain
concentrations as in equation (22). Figure 8 displays the
fluid velocities to visualize how fluids percolate through the
poroperm medium. Table 3 displays the flow results for
horizontal flow through the medium with both the
stress/pressure dependent inverse power rule of equation
(17) and with the shear strain concentration induced
fracture connectivity enhancment method. A 2 Mpa over
pressurization induces shear strain damage that increases
fluid throughput by about 10% from the zero stimulation
case (1.94 kg/s to 2.13 kg/s).
A curious result of the inverse power method is that the
inflow value is higher than the outflow. A possible physical
explanation for this phenomenon lies in the fact that the
flow calculations are made along the parallel vertical edges.
The left edge (inflow) has high pressurization of the
injection well acting on its surface to effectively reduce the
compressive stress and increase permeability, while the
right edge (outflow) has increased compressive stress acting
on it due to the the pressure from the left hand well. The
increased compressive stress acts to reduce permeability
and divert fluids away from the receiver well. The enhanced
permeability is no longer isotropic because the inverse
power relationship uses the full stress tensor, the individual
components of the permeability tensor can be affected, i.e.,
xx, xy, yy, and yx.
An encouraging result of the shear strain induced damage
relationship when applied to a poroperm medium is that
damage is high in areas where natural percolation pathways
exist. In Figure 7, notice the high shear strain
concentrations in the bottom right-hand portion of the
domain. The fluid velocity plot from Figure 8 shows that
the highly strained area is also where fluid velocity and
flow is the greatest. This area is then damaged the most and
flow is increased even more in the area with initially high
fracture connectivity. This is consistent with reported
results from the Rosemanowes site by Tester, et al. (2006).
They state that it is possible to stimulate natural fractures
for improved permeability. Figures 7 and 8 do show this
type of phenomenon. Further, this can easily lead to the
short-circuit case mentioned by Tester, et al. (2006) where
too much of the injection fluid passes through to the
receiver well too quickly without sufficient time to gather
heat energy.

Figure 7. Permeability enhancement by fracture


connectivity increases due to shear strain concentrations
in the medium. Fluid flows from left to right through
the medium. The left image displays shear strain
concentrations that result from fluid percolating
through native permeability pathways in the rock that
can be seen in Figure 8. The fracture connectivity
parameter was then increased in areas of high shear
strain concentrations. The center image displays the
initial permeability field and the right image shows the
enhanced
permeability
field.
The
maximum
permeability increased from 1.22E-13 to 1.33E-13 m2.

Figure 8. Fluid velocities in m/s for horizontal flow in


the poroperm medium. The left image displays the shear
strain concentrations, the center image displays the fluid
velocity through unenhanced pathways, and the right
image displays fluid velocity with strain-induced
enhancement in fracture connectivity.
Table 3. Comparison of inflow and outflow rates for
horizontal well to well flow with different stimulation
rules.

Permeability enhancement simulations were run on the


three spot well bore geometry seen in Figure 5 with the
poroperm medium description. The flow results for the
injection and receiver wells can be seen in Table 4.
Similarly, to Table 3, increases in compression around
receiver wells result in decreased permeability around those
wells. Hence, the slight reduction in outflow with the
inverse power rule for stimulation. Figure 9 displays the
shear strain concentrations around the injection well that
result from fluids percolating through natural permeability
channels in the poroperm medium. The results seem to
indicate that induced damage increases input well flow but
the effect does not carry to the output well. We learn from
this that (i) both input and output wells need stimulation,
and (ii) near-well fracture stimulation will have to be large
to permit stimulation effects away from the well.

New Zealand Geothermal Workshop 2012 Proceedings


19 - 21 November 2012
Auckland, New Zealand

Figure 9. Normalized shear strain concentrations


around a three spot wellbore geometry within a
poroperm medium. The shear strain concentrations lead
to increase in fracture connectivity, which in turn
increases permeability.
Table 4. Comparison of inflow and outflow rates for
three spot well bore simulations with different
stimulation rules.

5. CONCLUSIONS AND FUTURE WORK


It is clear that the physical processes that control
permeability evolution in EGS volumes are not known. Past
literature provides potential relations that could be used, but
it the physical basis of these relationships is unclear.
Further, past relationships, such as the inverse power rule,
do not naturally account for damage that may occur in the
medium. We propose a method that utilizes shear strain
concentrations where fracture damage and fracture
connectivity will increase. When applied to a poroperm
medium with realistic spatial fluctuations in porosity and
permeability, the strain induced damage method is able to
show fracture stimulation that is qualitatively similar to the
behavior reported at the Rosemanowes EGS site. Iis highly
desirable to obtain injection/receiver flow data and
porosity/permeability data from an EGS site, such as
Rosemanowes, to test the validity of the shear strain
induced damage model in a more quanitative way.
A major limiting factor however, is that the exact physics
and relationships that govern the permeability increase still
remain unclear. Further, our work has been limited to small
strain elastic behavior. It is imperitive to extend this work
to include irreversible deformations, such as plasticity
and/or damage. Figure 10 displays the Von Mises yield
criterion (J2) result for the three spot well bore problem
calculated as outlined in Chen and Han (1995). Areas of
high principal stress differentials would be subject to
irreversible deformations such as damage or plasticity. It
will be important to consider the effects that a
heterogenous, potentially, non-isotropic medium have on
the irreversible damage in future analyses.

Figure 10. Von Mises J2 yield criterion around the three


spot well bore setup. Areas that surpass a J2 yield
criterion could be subject to a different bulk constitutive
behavior that may include plasticity and/or damage.
ACKNOWLEDGEMENTS
The authors would like to acknowledge Dr. Rob Podgorney
of Idaho National Laboratories and David Dempsey of The
University of Auckland for their helpful discussions related
to this work.
REFERENCES
Bennion D and Griffiths J (1966). A Stochastic Model for
Predicting Variations in Reservoir Rock Properties.
Society of Petroleum Engineers Journal, pp. 9-16.
Chen, W.F. and Han, D.J. (1995) Plasticity for Structural
Engineers. Gau Lih Boook Co. Ltd.
Cladouhos, T., Petty, S., Larson, B., Iovenitti, J., Livesay,
B., and Baria, R. Toward More Efficient Heat
Mining: A Planned Enhanced Geothermal System
Demonstration Project. Geothermal Resources
Council Transactions, Vol 33. 2009.
Clauser, C. and Huenges, E. Thermal Conductivity in
Rocks and Minerals, Rock Physics and Phase
Relations, A Handbook of Physical Constants, AGU
Reference Shelf 3, 1995.
Goldstein, B.A., Hiriart, G., Tester, J., B., Bertani, R.,
Bromley, Gutierrez-Negrin, L.,C.J., Huenges, E., H,
Ragnarsson, A., Mongillo, M.A. Muraoka, and V. I.
Zui, (2011) Great expectations for geothermal energy
to 2100. Proceedings Thirty-Sixth Workshop on
Geothermal
Reservoir
Engineering,
Stanford
University
Hughes, T.J.R. (2000) The Finite Element Method: Linear
Static and Dynamic Finite Element Analysis. Dover
Publications.
Ingebritsen, S., Sanford, W., and Neuzil, C. (2006)
Groundwater in Geologial Processes. Cambridge
University Press.
Law, J. (1944) A Statistical Approach to the Interstitial
Heterogeneity of Sand Reservoirs Society of
Petroleum Engineers Journal. pp 202-222.
Leary, P.C. Fractures and physical heterogeneity in crustal
rock, in Heterogeneity of the Crust and Upper Mantle
Nature, Scaling and Seismic Properties, J. A. Goff,

New Zealand Geothermal Workshop 2012 Proceedings


19 - 21 November 2012
Auckland, New Zealand

& K. Holliger (eds.), Kluwer Academic/Plenum


Publishers, NewYork, 155-186, (2002) .
Leary, P.C., Pogacnik, J.A., Malin, P.E. (2012). Fractures
~ Porosity -> Connectivity ~ Permeability -> EGS
Flow Stimulation. Geothermal Resources Council
Transactions, Vol 36. 2012.
Leary, P.C. and Walter, L.A.: Crosswell Seismic
Applications to Highly Heterogenous Tight Gas
Reservoirs. First Break 26 (2008).
Lewis, R.W. and Schrefler, B.A. (1998). The Finite
Element Method in the Static and Dynamic
Deformation and Consolidation of Porous Media.
John Wiley and Sons. pp. 9 - 97, 341 - 394.
Likhachev, E.R. (2003). Dependence of Water Viscosity
on Temperature and Pressure. Technical Physics
48(4) 514-515.
Nathenson, Manuel (1999). The Dependence of
Permeability on Effective Stress from Flow Tests at Hot
Dry Rock Reservoirs at Rosemanowes (Cornwall) and
Fenton Hill (New Mexico) Geothermics 28. pp. 315340.

Pogacnik, J., Podgorney, R., and Leary, P. (2012).


Computational EGS: Modelling Fully-Coupled Heat
and Fluid Flow Through a Deformable FracturePermeable Medium with the Finite Element Method.
2012 Conference of the Engineering Mechanics
Institute. Notre Dame, Indiana.
Sutter, D., Fox, D.B., Anderson B.J., Koch. D.L., Rudolf
von Rohr, P., Tester, J.W. (2011), Sustainable heat
farming of geothermal systems: a case study of heat
extraction and thermal recovery in a model EGS
fractured reservoir. Proceedings Thirty-Sixth
Workshop on Geothermal Reservoir Engineering,
Stanford University.
Tester, et al. (2006). The Future of Geothermal Energy. A
Massachusetts Institute of Technology report.
http://geothermal.inel.gov/publications/future_of_geot
hermal_energy.pdf
Zienkiewicz OC & Taylor RL (2000) The Finite Element
Method: Volume 1 The Basis, 5th Ed. Butterworth
and Heinemann.

New Zealand Geothermal Workshop 2012 Proceedings


19 - 21 November 2012
Auckland, New Zealand

You might also like